1. Introduction
1.1. Invariant Subspace
The invariant subspace problem has been stated by Beurling and von Neumann [
1]. It can be formulated as follows.
Problem 1. Does every bounded linear operator on a given linear space have a non-trivial invariant subspace?
In 1966, Bernstein et al. [
2] showed that if
T is a bounded linear operator on a complex Hilbert space
and
p is a nonzero polynomial such that
is compact, then
T has non-trivial invariant subspace. Especially, when
, which is,
T itself is compact, the result was proved independently by von Neumann and N. Aronszajn, and in [
3], this result was extended to compact operators on a Banach space.
Let
T be a bounded linear operator on a Banach space. In 1973, Lomonosov [
4] proved that if
T is not a scalar multiple of the identity and commutes with a nonzero compact operator, then
T has a non-trivial hyperinvariant subspace, which is, any bounded linear operator commuting with
T has a non-trivial invariant subspace (other results see [
5,
6,
7]).
In 1976, Enflo [
8] was the first to construct an operator on a Banach space having no non-trivial invariant subspace and Nordgren et al. [
9] proved that every operator has an invariant subspace if and only if every pair of idempotents has a common invariant subspace.
In 1983, Atzmon [
10] constructed a nuclear Fréchet space
and a bounded linear operator, which has no non-trivial invariant subspace. Especially, in 1984, C. J. Read made an example, such that there is a bounded linear operator without non-trivial invariant subspace on
[
11].
In 2011, Argyros et al. [
12] constructed the first example of a Banach space for which every bounded linear operator on the space has the form
where
is a real scalar and
K is a compact operator, such that every bounded linear operator on the space has a non-trivial invariant subspace.
In 2013, Marcoux et al. [
13] showed that, if a closed algebra of operators on a Hilbert space has a non-trivial almost-invariant subspace, then it has a non-trivial invariant subspace (more results see [
14,
15,
16,
17,
18]).
In 2019, Tcaciuc [
19] proved that, for any bounded operator
T acting on an infinite-dimensional Banach space, there exists an operator
F of rank at most one such that
has an invariant subspace of infinite dimension and codimension.
For finite-dimensional vector spaces or nonseparable Hilbert spaces, the result is trivial. However, for infinite-dimensional separable Hilbert spaces, the problem is, after a long period of time, not yet completely solved.
1.2. Linear Dynamics
With the development of operator theory and dynamics progress, there are many papers about
C*-algebras and dynamics. Additionally, “the fundamental theorem of
C*-algebras [
20]” is Gelfand-Naimark theorem [
21]. Subsequently, in [
22], Fujimoto said that this theorem eventually opened the gate to the subject of
C*-algebras. Hence, there are various attempts to generalize this theorem [
23,
24,
25,
26,
27].
For the research on Problem 1 and with the development of chaos, Operator Dynamics or Linear Dynamics has aroused extensive attention as an important branch of functional analysis, which was probably born in 1982 with the Toronto Ph. D. thesis of C. Kitai [
28]. More details of this subject can be found in [
29,
30,
31,
32,
33].
If
X is a metric space and
T is a continuous self-map on
X, then the pair
is called a topological dynamic systems, which is induced by the iteration
Moreover, if T is a continuous invertible self-map on X, then is called an invertible dynamic and if the metric space X and the continuous self-map T are both linear, then the topological dynamic systems is called a linear dynamic.
For invertible dynamics, the relationship of Li-Yorke chaos between
and
was raised by Stockman as an open question [
34]. Additionally, in [
35,
36] and [
37], the authors give counterexamples for this question in noncompact spaces and compact spaces, respectively. For an invertible bounded linear operator
, the chaotic relationship between
and
is also interesting.
Next, we give the following definition
Definition 1 (Li-Yorke chaos)
. Let . If there exists , such that satisfies:then the operator T is said to be Li-Yorke chaotic, and x is called a Li-Yorke chaotic point of T. An example of an operator
T that is Li-Yorke chaotic but
is not can be found in [
38]. However, presently there is no general method to do this research. In fact, the
C*-algebra
generated by
T cannot be used for that.
1.3. Motivation and Main Results
For an
n-tuple
of not necessarily commuting operators, Colombo et al. [
39] put to use the notion of slice monogenic functions [
40] to define a new functional calculus, which is consistent with the Riesz–Dunford calculus in the case of a single operator and that allows the explicit construction of the eigenvalue equation for the
n-tuple
based on a new notion of spectrum for
(more results, see [
41,
42,
43,
44]).
In 2010, for bounded operators defined on quaternionic Banach spaces, Colombo et al. [
45] developed a noncommutative functional calculus that is based on the new notion of slice-regularity and that is based on the key tools of a new resolvent operator and a new eigenvalue problem, also, they extended this calculus to the unbounded case [
46] (more results, see [
47]).
In 2018, Monguzzi et al. [
48] characterized the closed invariant subspaces for the (*−) multiplier operator of the quaternionic space of slice
functions, obtained the inner-outer factorization theorem for the quaternionic Hardy space on the unit ball and provided a characterization of quaternionic outer functions in terms of cyclicity.
In this paper, we give a noncommutative functional calculus for . Additionally, by this construction, we give some applications, such as its applications on the invariant subspace problem and chaos. The precise meaning of the multiplication operator will become clear in Theorem 3.
Let
be a separable Hilbert space over
,
be the set of all bounded linear operator on
. For any given
, we obtain a
C*-algebra
associated with the polar decomposition
, where
and
is a
-cyclic vector, such that
. In this paper, we construct an isomorphism
, such that the following diagram is valid.
where
is the corresponding unitary operator associated with the isomorphism
and
,
and
. With this construction, we get a noncommutative functional calculus for the operator
T such that
Especially,
, which is the special case for normal operators, will become clear in Corollary 3, and, in this special case, we get that the noncommutative functional calculus of a normal operator
S is just only
, which is compatible with the classical normal operator functional calculus of [
49]. Where
and
.
Moreover, from , we deduce a sufficient condition to Problem 1 on infinite-dimensional separable Hilbert spaces and present the Lebesgue class , such that, if T is a Lebesgue operator, then T is Li-Yorke chaotic if and only if is.
In fact, we get that
and
where
is the set of all normal operator on
.
2. Decomposition and Isomorphic Representation
In this paper, means the conjugate of the complex function . Let X be a compact subset of , be the set of all continuous function on X, and be the set of all polynomial on X. For any given , let be its spectrum.
Following the polar decomposition theorem [
50] (p. 15), we get that
Let be the complex -algebra generated by and 1. Obviously, if T is invertible, then U is a unitary operator.
Lemma 1. Let be a compact subset not containing zero. If is dense in , then is also dense in .
Proof. By the properties of complex polynomials, we get that is a subalgebra of , which is closed under the standard algebraic operations. In addition, we have:
(1) ;
(2) separate the points of X;
(3) If , then .
We get the conclusion from the Stone–Weierstrass theorem [
49] (p. 145). □
For , there is . With Lemma 1, we get the following result.
Lemma 2. Let . If is dense in , then is also dense in .
Using the GNS construction [
49] (p. 250), for the
-algebra
, we have the following decomposition.
Lemma 3. Let T be an invertible bounded linear operator on . Then there exists a sequence of nonzero -invariant subspaces , such that:
(1) ;
(2) For every , there is a -cyclic vector such thatand Proof. The decomposition of (1) is obvious [
51] (p. 54), Therefore,
that is,
From Lemma 1, we get that
and
Let
be a
-cyclic vector, such that
is dense in
. Because the spectrum is closed and
, on
, we can define the nonzero linear functional
It is easy to get that
is a positive linear functional. By [
51] (p. 54), and the Riesz–Markov theorem, on
, we get that there is a uniquely finite positive Borel measure
, such that
Theorem 1. Let T be an invertible bounded linear operator on , be the complex -algebra generated by and 1 and let be a -cyclic vector, such that , where . Subsequently:
(1) there is a uniquely positive linear functional (2) there is a uniquely isomorphic representation associated with the uniquely finite positive Borel measure , which is complete.
Proof. (1) For
-cyclic vector
, we define the linear functional
We get that, on
, there is a uniquely finite positive Borel measure
, such that
Moreover, we can complete this Borel measure
on
. For this completion, we keep the notation
. We know that this Borel measure is unique [
52].
For any
, because of
we get that
is a positive linear functional.
(2) We know that
is dense in
. For any
, we get
Therefore,
is a surjective isometry from
to
.
Obviously,
and
are dense subspaces of
and
, respectively. Additionally, its closed extension
is an isomorphic operator.
Therefore, we get that is the uniquely isomorphic representation of associated with the uniquely finite positive Borel measure , which is complete. □
Let
T be an invertible bounded linear operator on
and
be a
-cyclic vector such that
. If there exists a unitary operator
, such that
, then
and we get two series of isomorphic representations
and
Let
Subsequently,
is a
-cyclic vector, such that
and we get the following equation
and
3. Noncommutative Functional Calculus
We know that the spectral theory and functional calculus of normal operators [
49] is very important in the study of operator theory and
-algebras [
50]. Inspired by the Hua Loo-kang theorem on the automorphisms of a sfield [
53], in this section, we give a useful construction from
to
and with this construction, we give a noncommutative functional calculus for any given
. However, there is valueless information just only from
or
.
Lemma 4. Let T be an invertible bounded linear operator on . Subsequently, we get Proof. That is, □
Let
T be an invertible bounded linear operator on
,
be a
-cyclic vector, such that
. On
, with
, we define the mapping
Following Lemma 3 and Theorem 1,
and
are dense subspaces of
and
, respectively. Its closed extension
is linear and for this closed extension we keep the notation
.
Subsequently, we obtain that
and
By a simple computation, we get that
By an application of the Banach inversion theorem [
49] (p. 91), we get that
is an invertible bounded linear operator from
to
.
Next, we define the operator
By Lemma 1 and [
51] (p. 55), we get that
is an invertible bounded linear operator on the Hilbert space
and
Moreover, we obtain the following diagram.
By [
53] and the isomorphic representations
and
of
, also, by Lemma 2 and 3, naturally, we give the following definition.
Definition 2. For invertible , let the symbol stand for . Subsequently, we get that there is a linear algebraic isomorphism from to , such that Let ξ be a -cyclic vector, such that and be a unitary operator such that . Subsequently, on we defineObviously, is dense in and is dense in . Then its closed extension isFor this closed extension, we keep the notation . With the polar decomposition theorem [
50] (p. 15), there is
. For invertible
, we get that
In fact, when T is invertible, we can choose a special unitary operator, which shows that the operators and are unitary equivalent. This is explained in the following theorem.
Theorem 2. Let T be an invertible bounded linear operator on and be a unitary operator, such that . Afterwards, there is a unitary operator , such that Moreover, is the corresponding unitary operator associated with the almost everywhere nonzero function , such thatwhere and ξ is a -cyclic vector, such that . Proof. By Lemma 3, let be a -cyclic vector, such that . By Definition 2, we have the linear operator
This construction yields that is an invertible linear operator from to . Hence, is an invertible linear operator from to .
By [
53], we get that
is a linear algebraic isomorphism from
on
to
on
. Additionally, by Lemma 2,
is dense in
and
is dense in
.
Hence, we obtain
that is,
and
are mutually absolutely continuous. Following [
49], (IX Theorem 3.6) and the construction
, we get that there exists
, where
, a.e., such that
From Lemma 4, for any
, because of
with [
50] (p. 60), we get that there is a unitary operator
, such that
That is,
is a
-cyclic vector. Additionally, with Theorem 1, we get
and
By a simple computation, we obtain that
and ≜ is introduced by
.
Hence, is an isomorphism from to .
With Theorem 1 and Definition 2, we have the operator
That is,
, such that the following diagram is valid.
Therefore, we said that the linear operator
is associated with
. Subsequently, we see that
is a unitary operator and by Lemma 3, we obtain
Afterwards,
is the corresponding unitary operator associated with
, which is, associated with the almost everywhere nonzero function
, such that
where
. □
We easily deduce
and the next results also readily follows. Let
T be an invertible bounded linear operator on
,
be a
-cyclic vector, such that
and let
be a unitary operator, such that
. In the proof of Theorem 2, and with the isomorphic representations
and
of
, we provide that
Especially, let
. Subsequently,
Corollary 1. Let T be an invertible bounded linear operator on and let ξ be a -cyclic vector, such that . Then and the equality is valid. Moreover, is the corresponding unitary operator associated with , whih is, associated with the almost everywhere nonzero function , such thatwhere . Next, for any given
, we define
Theorem 3. Let T be an invertible bounded linear operator on and be its Polar Decomposition. Let ξ be a -cyclic vector, such that and . Subsequently, there exists , such that and . Here .
Proof. Let
in the proof of Theorem 2. Afterwards, we get
By the polar decomposition theorem [
50] (p. 15) we have
. Hence, we get
By Corollary 1, we get
that is,
With the fact that
is a maximal abelian von Neumann algebra in
and the Fuglede–Putnam theorem [
49] (p. 279), we obtain that there exists
such that
Corollary 2. Let . Suppose and let ξ be a -cyclic vector, such that . Subsequently, there exists a function , such that Proof. For
,
is an invertible bounded linear operator on
. By the proof of Theorem 3, we get that
that is,
□
The following definition is quite natural.
Definition 3. For any given , we say that is the noncommutative functional calculus of T on , where ξ is a -cyclic vector, such that , and .
In the final part of this section, we give some properties of normal operator through the noncommutative functional calculus.
Corollary 3. For , if , then there exists such that is the noncommutative functional calculus of T on and we get . Where , ξ is a -cyclic vector, such that and Proof. For
and
, we get that
Therefore, we see that
and there exists
such that
That is,
With the proof of Theorem 3, we get , which is, □
Corollary 4. Let . Subsequently, the operator T is normal if and only if T is unitary equivalent to on , and if and only if , where ξ is a -cyclic vector, such that , , and .
4. A Sufficient Condition
In this section, we study Problem 1 on infinite-dimensional separable Hilbert spaces. With the fact that the exist of non-trivial invariant subspace is unchanged by the similarity of bounded linear operators on Banach spaces [
1], which is, for
and
, if
is an invertible bounded linear operator and
, then
R has non-trivial invariant subspace if and only if
S has, where
and
are Banach spaces. Therefore, for any given
, using the construction of
, we give a sufficient condition to Problem 1 on infinite-dimensional separable Hilbert spaces.
For convenience, we define . Obviously, is a closed subspace of .
Theorem 4. Let , and be the noncommutative functional calculus of on where , ξ is a -cyclic vector, such that and . If , then T has a non-trivial invariant subspace.
Proof. It is enough to prove the result for infinite-dimensional separable complex Hilbert space . Obviously, if is a non-trivial invariant subspace of T if and only if A is a non-trivial invariant subspace of , where .
Let
and let
be a
-cyclic vector such that
. Subsequently, following Corollary 2, we get that there exists
, such that
is the noncommutative functional calculus of
on
By the construction of in Theorem 2, we get and .
(1) If
, that is
, by the proof of Corollary 3, then
T is unitary equivalent to
. Because
is a normal operator, it possesses a non-trivial invariant subspace and, hence, the same is true for
T. For details, see, e.g., [
49].
(2) If
and
, then
is a non-trivial invariant subspace of
and we get
Hence,
is a non-trivial invariant subspace of
. With the proof of Theorem 3, we get that
5. Lebesgue Operator
In this section, we study chaos of an invertible bounded linear operator on an infinite-dimensional separable Hilbert space. For the example of integral calculus in mathematical analysis, we know that the convergence or the divergence of the weighted integral calculus of x and should be independent of each other; however, sometimes it happens that this indeed depends on a special choice of the weight function.
In the view of integral calculus, we define the Lebesgue class and prove that if T is a Lebesgue operator, then T is Li-Yorke chaotic if and only if is. With the idea of the noncommutative functional calculus , we give an example of a Lebesgue operator that is not a normal operator.
Let
be the Lebesgue measure on
. By Theorem 1, there exists a Borel measure
, which is complete, such that
is a Hilbert space. If there exists
, such that, for all
, the measure
is absolutely continuity with respect to
, then using the Radon–Nikodym theorem [
49] (p. 380), there exists
, such that
, where
,
and
.
Definition 4. Let T be an invertible bounded linear operator on the separable Hilbert space over . Suppose that the operator T satisfies the following conditions:
There exists , such that, for all There exists , such that for all and for any given nonzero , there exists a nonzero function and a nonzero vector , such that whenever .
Subsequently, the operator T is said to be a Lebesgue operator, and the family of all Lebesgue operators on is denoted by .
Theorem 5. Let T be a Lebesgue operator on the separable Hilbert space over . Subsequently, T is Li-Yorke chaotic if and only if is.
Proof. Let
be a
-cyclic vector such that
If
is a Li-Yorke chaotic point of
T, then by Definition 4, we see that, for
large enough, there exist
,
and
, such that
,
, and
Therefore, we get the following
where ≜ is following Definition 4. By Definition 1, we get that
T is Li-Yorke chaotic if and only if
is. □
Following [
54], for
and
, we introduce the distributional function
In addition, we denote
and introduce the following definition.
Definition 5. Let . If there exists and
(1) If , for some and for , then we say that T is distributionally chaotic or I-distributionally chaotic.
(2) If for and for , then we say that T is -distributionally chaotic.
(3) If for , then we say that T is -distributionally chaotic.
Corollary 5. Let T be a Lebesgue operator on the separable Hilbert space over . Then T is I-distributionally chaotic (or -distributionally chaotic or -distributionally chaotic) if and only if is I-distributionally chaotic (or -distributionally chaotic or -distributionally chaotic).
Theorem 6. There exists an invertible bounded linear operator T on the separable Hilbert space over , such that T is Lebesgue operator that is not a normal operator.
Proof. Let . Subsequently, is a separable Hilbert space over . Any separable Hilbert space over can be expanded to a separable Hilbert space over . Without loss of generality, let be the separable Hilbert space over . We prove the conclusion by six parts:
(1) Let
. We construct a measure preserving transformation on
. Let
. We get a Borel algebra
generated by
M. We define
,
Subsequently,
is an invertible measure preserving transformation on the Borel algebra
. With [
55] (p. 63),
and
is a unitary operator associated with
, where
is the operation of composition
(2) Define on . Subsequently, is an invertible positive operator.
(3) For , , we define . Afterwards, is continuous and , a.e., . Hence, that is absolutely continuous with respect to is a finite positive Borel measure that is complete. That is, is a separable Hilbert space over . Moreover, and are unitary equivalence.
(4) Let
. We get
Because of
we get that
T is not a normal operator and
.
(5) Let the operator on be corresponding to the operator on . Subsequently, is an invertible bounded linear operator that is not a normal operator and .
We get that is continuous and almost everywhere positive. Hence, is a finite positive Borel measure that is complete.
For any
, we define
when
else
. Subsequently,
is the identity function on
E. With a simple computing, we get that
and
We see that
. From
and
let
.
Afterwards, is a finite positive Borel measure that is complete. For any given nonzero , we get the nonzero function .
Easily, we get that is a -cyclic vector of the multiplication and is a -cyclic vector of . By Definition 4, we get that is Lebesgus operator, but is not a normal operator. □
Corollary 6. There exists an invertible bounded linear operator T on the separable Hilbert space over , such that T is a Lebesgue operator that is a positive operator.
Corollary 7. Let be the subspace of all normal bounded linear operator on an infinite-dimensional separable Hilbert space . Subsequently, the following families of linear operators are non-empty:In fact, both these families contain non-trivial members.